You are on page 1of 14

Journal of Volcanology and Geothermal Research 369 (2019) 21–34

Contents lists available at ScienceDirect

Journal of Volcanology and Geothermal Research

journal homepage: www.elsevier.com/locate/jvolgeores

Eruption and fountaining dynamics of selected 1985–1986 high


fountaining episodes at Kīlauea volcano, Hawai'i, from quantitative
vesicle microtexture analysis
S.J. Holt a,⁎, R.J. Carey a, B.F. Houghton b, T. Orr c, J. McPhie a, S. Feig a
a
School of Natural Sciences and CODES, University of Tasmania, Private Bag 79, Hobart 7001, Tasmania, Australia
b
Department of Geology and Geophysics, University of Hawai'i, Honolulu, HI 96822, USA
c
U.S. Geological Survey, Alaska Volcano Observatory, 4230 University Drive, Suite 100, Anchorage, AK 99508, USA

a r t i c l e i n f o a b s t r a c t

Article history: Tephra from the early Hawaiian fountaining episodes of the ongoing eruption of Pu'u 'Ō'ō in the East Rift Zone
Received 11 December 2017 (ERZ) of Kīlauea provides an opportunity to study the vesicle microtextures of pyroclasts erupted from a single
Received in revised form 15 November 2018 vent over a prolonged period of time. We report the results of microtextural analysis of pyroclasts from five of
Accepted 17 November 2018
Pu'u 'Ō'ō's high (N200 m) Hawaiian fountaining episodes (episodes 32, 37, 40, 44 and 45) erupted during
Available online 19 November 2018
1985–1986. This analysis was carried out to constrain the parameters that led to large variations in fountain
Keywords:
height at Pu'u 'Ō'o, and the extent to which pyroclast residence times in the fountain modified microtextures.
Kīlauea Our results confirm the finding of Stovall et al. (2011, 2012) that pyroclasts from a single Hawaiian fountain
Hawaiian fountaining can vary greatly in texture (from bubbly to foamy), and have vesicle volume densities (Nmv) and vesicle to
Pu'u 'Ō'ō melt ratios (VG/VL) that vary by an order of magnitude. This range in vesicle texture and population is due to ex-
Vesicle microtexture tensive growth and coalescence of vesicles within the fountain after fragmentation. Only one pyroclast from four
of five episodes was found to have textures interpreted as indicative of the vesicle population near the moment of
fragmentation: bubbly texture, high density (typically N500 kg m−3), high Nmv (2.2 × 106 to 4.4 × 106), and low
VG/VL of 2.06 to 4.65. We demonstrate a linear correlation between Δ(VG/VL) and peak fountain height across a
range of Hawaiian fountains from Kilauea. This correlation could be used to infer peak heights of unobserved
Hawaiian fountaining eruptions after further testing using well-recorded events.
© 2018 Elsevier B.V. All rights reserved.

1. Introduction Hawaiian fountaining is defined as the sustained jetting of gas and


fluidal coarse pyroclasts (Taddeucci et al., 2015; Wolff and Sumner,
The dynamic processes within magma during its ascent in the shal- 2000) and is a common phenomenon at basaltic volcanoes (Fedotov
low conduit of a volcano have a major influence on subsequent eruptive et al., 1980; Heliker et al., 2003; Polacci et al., 2006). Kīlauea Volcano
behaviour (Gonnermann and Manga, 2013; Houghton et al., 2004; is the type-locality for Hawaiian fountaining (Fig. 1). In its written his-
Houghton et al., 1999; Papale, 1998; Parfitt, 2004; Vergniolle and torical period (since 1823), Kīlauea has had three well-documented
Jaupart, 1990). Microtextural analysis of pyroclasts is used to quantify eruptions that produced multiple Hawaiian fountaining episodes: at
rates and timing of vesiculation processes in the conduit, and Kīlauea Iki in 1959 (Richter et al., 1970; Stovall et al., 2011; Stovall
fountaining dynamics during eruptions (Adams et al., 2006; Carey et al., 2012), at Mauna Ulu in 1969 (Parcheta et al., 2013; Swanson
et al., 2012; Costantini et al., 2010; Lautze and Houghton, 2007; et al., 1979) and at Pu'u 'Ō'ō from 1983 to 1986 (Heliker et al., 2003;
Parcheta et al., 2013; Porritt et al., 2012; Rust and Cashman, 2011; Wolfe et al., 1987).
Sable et al., 2006; Stovall et al., 2011; Stovall et al., 2012). Microtextural Hawaiian fountains constitute a low-energy end-member of
datasets can be used to quantify processes during a single episode or ex- sustained (long-duration), explosive volcanism (Houghton et al.,
plosion (Carey et al., 2012; Sable et al., 2006) and also to describe the 2016; Walker, 1973). Traditionally, the preferred model for the vesicu-
range of eruptive behaviour of a volcano during multiple eruption epi- lation style and dynamics of Hawaiian-style volcanism was based on
sodes over an extended period of time (Stovall et al., 2011; Stovall rapid ascent of the magma within the shallow conduit, permitting me-
et al., 2012). chanical coupling of the shallowest-exsolved and smallest bubbles of
magmatic volatile gases with the liquid melt (Cashman and Mangan,
⁎ Corresponding author. 1994; Parfitt and Wilson, 1995; Parfitt, 2004; Wilson, 1980; Wilson
E-mail address: sjholt@utas.edu.au (S.J. Holt). and Head, 1981). However, a lack of correlation between the mass

https://doi.org/10.1016/j.jvolgeores.2018.11.011
0377-0273/© 2018 Elsevier B.V. All rights reserved.
22 S.J. Holt et al. / Journal of Volcanology and Geothermal Research 369 (2019) 21–34

Fig. 1. A) Digital Elevation Model (DEM) of Island of Hawai'i. B) DEM showing the summit and the approximate location of the upper part of the East Rift Zone of Kīlauea Volcano.

discharge rate and eruption style, as well as observed rapid fluctuations (Edmonds and Gerlach, 2007; Spilliaert et al., 2006), bulk rheology
in the eruption intensity during a single episode, show that eruptive be- (Llewellin and Manga, 2005; Manga et al., 1998) and the rates of bubble
haviour is dependent not just on the rise rate of the magma, but also on nucleation, growth, coalescence and loss (Adams et al., 2006; Cashman
the depth, timing, and rate of volatile exsolution, and the buoyant ascent and Mangan, 1994; Lautze and Houghton, 2007; Sable et al., 2006) are
of large vesicles within the shallow conduit (Gonnermann and Manga, complex and interrelated, and govern the explosive eruption styles of
2007, 2013; Houghton and Gonnermann, 2008; Taddeucci et al., basaltic volcanoes. It is possible to quantify aspects of the vesiculation
2015). Variables including (but not limited to) the pre-eruptive volatile history, i.e., the rates of bubble nucleation, growth, coalescence and es-
content (Johnson et al., 1994; Papale, 2005), the depth of degassing cape in the magma within the shallow conduit (both pre- and post-

Fig. 2. Photographs of high Hawaiian fountains at Pu'u 'Ō'ō. 1) Incandescent, molten jet; 2) black ash and lapilli in the cooler outer region of the jet. A) Episode 45; view to the south. Peak
fountain height of episode: 257 m. Approximate fountain height at the time of photo: 200–250 m. B) Episode 44; view to the southwest. Peak fountain height of episode: 308 m. Fountain
height at the time of photo: 250–275 m.
S.J. Holt et al. / Journal of Volcanology and Geothermal Research 369 (2019) 21–34 23

Table 1
Eruption statistics from Heliker et al. (2003) for episodes 32, 37, 40, 44 and 45 of Pu'u 'Ō'ō eruption.

Episode Sample Episode start Episode end Repose periodb Episode length Area covered by lava DRE volumec (106 Peak fountain
number (date –timea) (date –timea) (days) (days) (km2) m3) height (m)

32 04B 21/04/85 – 1516 22/04/85 – 0906 38.4 0.8 4.8 11.4 391
37 04A 24/09/85 – 1808 25/09/85 – 0619 21.8 0.5 4.4 10.3 352
40 03 01/01/86 – 1309 02/01/86 – 0238 48.5 0.6 4 8.1 264
44 02 13/04/86 – 2054 14/04/86 – 0756 22.2 0.5 5.2 8.1 308
45 01 07/05/86 – 2241 08/05/86 – 1106 23.6 0.5 5.5 6.6 257

Data from Heliker et al. (2003).


a
Hawaii-Aleutian Time Zone, UTC/GMT-10.
b
Repose period leading up to episode.
c
Dense rock equivalent volume.

fragmentation), through microtextural analysis of the erupted tephra This study compares five of the later high Hawaiian fountaining ep-
(Adams et al., 2006; Cashman and Mangan, 1994; Lautze and isodes from Pu'u 'Ō'ō in the East Rift Zone (ERZ) in 1985–1986, which
Houghton, 2005; Mangan and Cashman, 1996; Parcheta et al., 2013; erupted magma of nearly constant composition but had varying foun-
Polacci, 2005; Sable et al., 2006; Stovall et al., 2011; Stovall et al., 2012). tain heights and durations. We quantify the vesiculation histories
The quantification of vesicularity has also been used to identify the through microtextural analysis of tephra erupted during episodes 32,
locations, cooling rates and residence times of pyroclasts within Hawai- 37, 40, 44 and 45. The microtextural characteristics of the deposits are
ian fountains (Porritt et al., 2012). Stovall et al. (2011, 2012) reported on then linked to visual observations of fountain geometries made by the
Hawaiian pyroclast morphologies coupled with qualitative and quanti- staff of the U.S. Geological Survey's Hawaiian Volcano Observatory
tative microtextural evidence to show the progression of post- (HVO). The data are used to answer three key questions: (1) do
fragmentation expansion during transport time within a fountain. How- lapilli-sized pyroclasts from these episodes retain any signature of
ever, these authors were able to demonstrate that clast rims with a high syn-eruption magma properties? (2) does fountain height (reflecting
vesicle number density are a valid approximation to magma conditions decompression rate) scale with bubble number density of lapilli? and
pre-fragmentation. Further, Stovall et al. (2012) could demonstrate that (3) is there any quantifiable textural signature (such as vesicle-to-
clasts with the highest vesicle number densities equate to the highest melt ratio, VG/VL) that can be used as a proxy for clast residence times,
intensity phase, Episode 15. and therefore fountain height, for Pu'u 'Ō'ō fountains?

Fig. 3. A) Stratigraphic column of the deposit sampled, showing eruption episode number and sample set number for each unit. B) The location of the sample site relative to the Pu'u 'Ō'ō
vent. The photo was taken after episode 28 on December 15, 1984. Photo courtesy of HVO.
24 S.J. Holt et al. / Journal of Volcanology and Geothermal Research 369 (2019) 21–34

2. Eruption history of Pu'u 'Ō'ō vesicularity for each pyroclast was then calculated using a DRE density
of 2950 kg m−3 following Ryan (1988). Three or four pyroclasts were
The Pu'u 'Ō'ō eruption began on the 3rd of January 1983. After three chosen from each sample for microtextural image analysis. The choice
months dominated by distributed fissure fountaining, activity localised of pyroclast was a function of the density distribution, with an emphasis
at a vent later named Pu'u 'Ō'ō (Fig. 1) (Heliker et al., 2003; Wolfe to select two pyroclasts that represent the bulk of the magma erupted.
et al., 1988). From June 1983 to June 1986, 44 Hawaiian fountaining ep-
isodes took place at Pu'u 'Ō'ō. The activity was highly cyclic during this
time period, with lengthy repose periods (8 to 65 days) punctuated by
relatively short (0.3 to 16 days) Hawaiian fountaining episodes. Repose
periods were characterised by fluctuating magma level within the con-
duit and outgassing from the vent during gradual inflation of Kīlauea's
summit. Fountaining episodes consisted of low (≤200 m) to high
(N200 m) Hawaiian fountaining and lava effusion (both pāhoehoe and
'a'ā lavas), accompanied by rapid deflation at the summit of Kīlauea
(Heliker et al., 2003; Wolfe et al., 1987; Wolfe et al., 1988). This pattern
was broken in July 1986, when the eruption shifted to a new vent
(Kupaianaha), and the eruptive style changed to one of nearly continu-
ous effusion (Heliker et al., 2003).

2.1. Studied eruption episodes

The five episodes discussed in this paper (episodes 32, 37, 40, 44 and
45) took place between April 1985 and May 1986. These episodes were
chosen because all fountains exceeded 200 m in height (e.g., Fig. 2), had
products that could be precisely linked to an episode and sampled, and
were documented in detail during eruption. The visual observations are
summarised in Table 1, modified from Heliker et al. (2003). Together,
these episodes erupted a total of 4.5 × 107 m3 (Dense Rock Equivalent,
DRE) of magma and had fountain heights ranging from 264 to 391 m.
In each episode, the maximum fountain height was reached after simul-
taneous gradual build up in intensity and summit deflation. At the end
of each episode, seismic tremor within the ERZ decreased rapidly and
summit deflation switched back to inflation (Heliker et al., 2003). The
geochemical composition of the magma became steadily more mafic
through the five episodes studied and contained no traces of the early
hybrid magma that was erupted at Pu'u 'Ō'ō until Episode 30 (Garcia
et al., 1992; Moore et al., 1980).

3. Sampling and analytical methods

3.1. Field work and sample selection

A suite of samples was collected on the 5th of March 2008 from a 56-
cm-thick deposit approximately 1 km north of Pu'u 'Ō'ō that included
beds corresponding to six eruptive episodes (Fig. 3). At the time of sam-
ple collection, this location was one of very few where parts of the
1983–1986 pyroclastic succession was accessible. However, it was over-
run by lava by 2013–2016. At this location, the stratigraphy of the
tephra was described, and samples of 100 clasts in a restricted size frac-
tion (16 to 32 mm in diameter) were collected from narrow vertical in-
tervals within each fall unit following the methodology of Houghton
and Wilson (1989). These samples were used for componentry analysis,
juvenile clast density, and more detailed microtextural work. Tephra
units from most of the Pu'u 'Ō'ō episodes are not present at this site,
as the dominant direction of pyroclastic fall was southwest due to the
prevailing trade winds. Because of the infrequency of tephra fall north
of the vent, it was therefore possible to identify the episodes sampled
with a high degree of certainty (Appendix A). The deposit of Episode
30 was not studied as we were unable to obtain a sample wholly repre-
sentative of the deposit.

3.2. Analytical methods

Fig. 4. Density histograms for the pyroclast sample sets from each episode. Small numbers
The bulk density of every pyroclast within each sample was mea- within the histograms correspond to the specific pyroclasts (listed on right) chosen for
sured following the method of Houghton and Wilson (1989), measure- microtextural analysis. For comparison, blue lines are at fixed densities of 400, 800 and
ment have a ±30 kg·m−3 error, and histograms were generated. Bulk 1000 kg·m−3.
S.J. Holt et al. / Journal of Volcanology and Geothermal Research 369 (2019) 21–34 25

Table 2
– Texture and vesicularity data for all analysed pyroclasts.

Pyroclast Episode Texturea Density (kg m−3) Density-derived vesic. (%) Nv Nmv VG/VLb

01-33 Ep. 45 B 960 67.3 5.2 × 105 1.6 × 106 2.06


01-08 F/B 840 71.4 6.1 × 105 2.2 × 106 2.50
01-06 F 620 79.0 1.5 × 105 7.3 × 105 3.76
01-01 F 460 84.6 3.9 × 104 2.5 × 105 5.49
02-04 Ep. 44 B 700 76.2 1.0 × 106 4.4 × 106 3.20
02-08 F 450 84.7 9.4 × 104 6.1 × 105 5.54
02-05 F/B 330 88.8 7.7 × 104 6.8 × 105 7.93
03-05 Ep. 40 B 770 74.1 8.9 × 105 3.5 × 106 2.86
03-07 B 520 82.3 6.8 × 105 3.8 × 106 4.65
03-17 F/B 360 87.8 7.9 × 105 6.5 × 106 7.20
03-08 F 260 91.2 3.1 × 104 3.5 × 105 10.36
04A-05 Ep. 37 B 940 68.3 1.3 × 106 4.2 × 106 2.15
04A-09 B 750 74.6 5.7 × 105 2.2 × 106 2.94
04A-01 F 330 88.7 3.7 × 104 3.2 × 105 7.85
04A-04 F 240 91.7 4.5 × 104 5.4 × 105 11.05
04B-20 Ep. 32 B 1120 62.1 3.3 × 105 8.9 × 105 1.64
04B-38 IB 870 70.7 1.7 × 105 5.7 × 105 2.41
04B-16 F 350 88.3 5.6 × 104 4.8 × 105 7.55
04B-06 F 290 90.2 5.1 × 104 5.3 × 105 9.20
a
F = foamy, B = bubbly, F/B = combination of both, IB = irregular bubbly.
b
Vesicle to melt ratio.

Additional pyroclasts were subsequently chosen near minimum and sets for each pyroclast were generated using both a Dell V105 scanner
maximum densities for samples with wide density distributions. Thin- and FEI Quanta 600 MLA Scanning Electron Microscope. Images were
sections were made from the selected pyroclasts and nested image taken at 4.5 cm, 5.0 mm and 1.25 mm horizontal field widths

Fig. 5. A) Binary image showing bubbly texture. B) Binary image showing foamy texture. C) Example of a pyroclast that exhibits both bubbly and foamy texture. Note that the transition
between the two textural domains is sharp. D) Binary image showing irregular bubbly texture.
26 S.J. Holt et al. / Journal of Volcanology and Geothermal Research 369 (2019) 21–34

Fig. 6. Four separate 500× magnification images showing the microlite species, abundance and textures from pyroclast 04B-38. In all images cpx: clinopyroxene, P: plagioclase, mag:
magnetite and ves: vesicle. See supplementary material for locations of images within the pyroclast.

(corresponding to 25×, 54× and 215× magnification) and nested to en- density distributions; modal densities are between 300 and
sure that full ranges of vesicle sizes were effectively imaged (Shea et al., 500 kg m−3, corresponding to vesicularities of 89% to 82%. Of these
2010). Each nest contained from seven to twenty-one images. Adobe three episodes, the density distributions of episode 37 and episode
Photoshop CS6 software was used to generate binary images which 44 pyroclasts are unimodal, whereas the episode 32 data are bimodal
were processed using the ImageJ software package (Abràmoff et al., and have a secondary mode at 700 to 900 kg m−3, corresponding to
2004) to measure the number of vesicles per unit area and their individ- vesicularities of 74% to 67%. Episode 32 has a much larger range of
ual size and shape. These measurements were converted to three di- pyroclast densities (70–1330 kg m−3) than episodes 37 and 44
mensions (vesicle volume density, Nv) using stereological techniques (240–1170 kg m−3 and 110–970 kg m−3 respectively). Although all
(Sahagian and Proussevitch, 1998). To account for the presence of vesi- three sample sets are positively skewed toward low density/high ve-
cles, a melt correction is applied to the Nv values to obtain the vesicle sicularity, the episode 32 pyroclasts have a much larger sub-
volume density per unit melt, Nmv (Klug et al., 2002). The vesicle-to- population of pyroclasts with densities N800 kg m −3 (b71%
melt ratio (VG/VL) was calculated for each representative pyroclast vesicularity).
using the following equation from Gardner et al. (1996): Episode 40 pyroclasts have a modal density similar to those of epi-
  sodes 32, 37 and 44 (300–400 kg m−3 and 82–89% vesicularity) but a
VG Ves 1 much flatter density distribution, whereby pyroclasts with densities be-
¼
V L 1−Ves 1−m tween 200 and 700 kg m−3 (74% to 93% vesicularity) are of more equal
abundance. Episode 40 pyroclasts also have the narrowest density range
where VG is the volume of gas (vesicles), VL is the volume of melt (glass), of the five episodes, all 100 pyroclasts having densities between 200 and
Ves is the volume fraction of vesicles and m is the modal fraction of 900 kg m−3 and vesicularities between 92% and 67%.
crystals. Episode 45 is the only episode that produced pyroclasts with a neg-
A detailed description of the image processing and microtextural atively skewed density distribution. The modal density is between 800
analysis methodology used can be found in Adams et al. (2006), Shea and 900 kg m−3 (70–67% vesicularity) and less well defined than for
et al. (2010), Stovall et al. (2011), and Parcheta et al. (2013). the three episodes that are strongly positively skewed (episodes 32,
37, 44). In addition to the highest modal density of all five episodes, ep-
4. Results isode 45 pyroclasts have the largest range, from 50 to 1350 kg m−3, cor-
responding to vesicularities between 98% and 50%. While episode 45
4.1. Juvenile clast density and vesicularity pyroclasts have densities mainly between 600 and 1000 kg cm−3,
there is a small secondary peak between 400 and 500 kg m−3, corre-
Fig. 4 shows the density distributions for the pyroclast sample sponding closely to the modal densities and vesicularities of the other
sets. Three episodes, episode 32, 37 and 44 have positively skewed four episodes.
S.J. Holt et al. / Journal of Volcanology and Geothermal Research 369 (2019) 21–34 27

04B-38 has a texture similar to bubbly texture, in that the vesicles are
separated by glass walls 100–250 μm thick. However, this pyroclast
also has relatively large (1.00 and 1.24 mm) vesicles that are commonly
polylobate or irregular in shape (Fig. 5D). Pyroclast 04B-38 is the only
one within the studied sample set that contains observable microlites
(Fig. 6).

4.2.2. Quantitative results


The vesicle number densities of the measured pyroclasts (Nv; num-
ber of vesicles per unit volume of bulk rock) vary by almost two orders
of magnitude across all the episodes sampled, from 3.0 × 104 cm−3 in
episode 40 (pyroclast 03–08) to 1.3 × 106 cm−3 in episode 37 (pyroclast
04A-05) (Table 2). Clasts representing a single episode have Nv values
that span at least one order of magnitude; the largest range is in episode
37 pyroclasts, with a minimum of 3.7 × 104 cm−3 and a maximum of 1.3
× 106 cm−3. Using melt-corrected Nmv the range in vesicle number den-
sities across all the episodes is from 2.5 × 105 cm−3 in episode 45
(pyroclast 01–01) to 6.5 × 106 cm−3 in episode 40 (pyroclast 03–17)
(Table 2). The Nmv values for these five Hawaiian fountaining episodes
at Pu'u 'Ō'ō are typically less than those calculated for the 1959 eruption
of Kīlauea Iki (Stovall et al., 2011; Stovall et al., 2012) and larger than
those calculated for the 1969 eruption of Mauna Ulu (Parcheta et al.,
2013) (Fig. 7).
To better define the vesiculation histories during each eruptive epi-
sode, vesicle size distributions (VSDs) were constructed for each
analysed pyroclast (Fig. 8). These graphs show the fraction of the total
vesicle volume that is attributed to each vesicle size range, expressed
as equivalent diameter, the diameter of a circle with the same area as
calculated for the vesicle (Adams et al., 2006). In most cases the VSDs
are unimodal and have a moderate range (0.04–1.24 mm) (Fig. 8).
Within two of the five eruption episodes (episodes 45 and 37), there
is a common trend of increasing unimodality and negative skew with
Fig. 7. Plot showing vesicularity versus vesicle number density (Nmv) for three high respect to increasing pyroclast vesicularity for VSDs. This trend is de-
Hawaiian fountaining eruptions at Kīlauea. This results in a b2% variation of vesicularity. fined by the migration of the median and modal vesicle size to larger
sizes, coupled with a decrease in the number of vesicles ≤0.16 mm in
4.2. Microtextural analysis size. The VSDs of episode 32 pyroclasts have a unique trend in that the
pyroclasts have a consistent modal vesicle size between 0.4 and
4.2.1. Qualitative results 0.6 mm. However, the proportion of vesicles ≤0.16 mm and ≥ 1.0 mm
The pyroclasts from all five eruption episodes have vesicle decreases and the proportion of vesicles between 0.5 and 0.6 mm in-
microtextures that can be classified into three distinct categories. The creases, so that the modal vesicle size becomes more dominant and
texture within each studied pyroclast is indicated in Table 2. The ‘bubbly the size range narrows with increasing vesicularity (Fig. 8).
texture’ category (‘B’ in Table 2) is dominated by small, spherical vesi-
cles (100–200 μm in diameter) that are separated by thick glass walls 5. Discussion
(commonly 50 μm thick, but up to 100 μm) (Fig. 5A). The round shape
suggests that there has been minimal interaction between neighbouring 5.1. Determining syn-eruptive vesicle textures
bubbles. Clasts with this texture are typically the least vesicular and
most dense. Within the incandescent fountain, the cooling rates of lapilli-
The ‘foamy texture’ category (‘F’ in Table 2) is dominated by large, sized clasts are predicted to be too slow to retain a signature of the
closely packed polygonal vesicles, typically ≥500 μm, with bubble magma properties at the time of fragmentation (Porritt et al.,
walls approximately 5 μm thick (Fig. 5B). The texture of these pyroclasts 2012). The residence time within the incandescent fountain results
closely resembles reticulite texture as defined by Mangan and Cashman in a textural evolution from a relatively dense fluidal clast to pumice,
(1996), whereby the vesicles are polyhedral and are separated by followed by quenching during transport in a fountain (Stovall et al.,
micrometer-scale glass septa. Clasts with this texture have very high ve- 2011, 2012). Stovall et al. (2011, 2012) showed that vesiculation
sicularities, but very few fall within the vesicularity range for reticulite, stages can be identified by comparing the Nmv of each pyroclast to
defined as N98% by Mangan and Cashman (1996), and therefore we use its vesicle-to-melt ratio (VG/VL) as defined by Gardner et al. (1996).
the name foamy for this textural end-member. Each separate vesiculation process, such as nucleation, growth, coa-
Although these two end-member textures differ greatly, they are not lescence and loss (and any feasible combination of these) produces
mutually exclusive, and there is a small population of pyroclasts that ex- a characteristic trend in both VG/VL versus Nmv plots and the associ-
hibit both textures, denoted as F/B in Table 2 (Fig. 5C). In such pyroclasts ated VSDs (Shea et al., 2010; Stovall et al., 2011; Stovall et al., 2012).
(only 3 of the total 19 in this study), the transition between bubbly and Following Stovall et al. (2011), we interpret pyroclasts that have rel-
foamy domains is abrupt without transitional regions. Furthermore, atively high Nmv and low VG/VL to resemble most closely the vesicu-
there is no consistent spatial distribution of the two end-member tex- lation conditions of the magma at fragmentation; we call these
tures in a pyroclast; for example, one end-member is not confined to ‘primary’ pyroclasts. For all five episodes, these primary pyroclasts
the rim or core of a pyroclast. (hollow data points in Fig. 9A–C), which we interpret to approximate
A third vesicle microtexture, which we call ‘irregular bubbly texture’ the syn-fragmentation vesicle population, have bubbly textures, low
(‘IB’ in Table 2), was found in one pyroclast from episode 32. Pyroclast VG /VL values and high Nm v (Fig. 9), and are likely to have been
28
S.J. Holt et al. / Journal of Volcanology and Geothermal Research 369 (2019) 21–34
Fig. 8. Histograms showing vesicle size distributions for representative pyroclasts from each high fountaining episode studied. Equivalent diameter is defined as the diameter of a circle with the same area as calculated for the vesicle (Adams et al.,
2006). Pyroclasts from each episode are arranged from top to bottom in order of increasing vesicularity (decreasing density). In each plot, the pyroclast number is given in the top right and the Nmv (bubbles cm−3), vesicularity (%) and density
(kg m−3) are listed in the top left. The vertical blue line through the 0.16 mm equivalent diameter is for comparison.
S.J. Holt et al. / Journal of Volcanology and Geothermal Research 369 (2019) 21–34 29

Fig. 9. VG/VL versus Nmv plots for each eruption episode showing the post-fragmentation vesiculation process that is defined by the pyroclast microtextures (after Stovall et al., 2011). In all
diagrams, open circles denote primary pyroclasts and the grey arrows highlight the associated trend. The images show the microtexture of the labelled pyroclasts that are representative of
textures found for that episode. A) Episodes 37, 44 and 45, all of which show growth plus coalescence trends. B) Episode 40, which shows both a growth plus coalescence trend and a
growth plus nucleation trend. C) Episode 32, which has a growth trend. D) Key showing the possible trends and their associated post fragmentation vesicle processes. L = loss, C =
coalescence, G + C = growth plus coalescence, G = growth, G + N = growth plus nucleation, N = nucleation.

transported in the cooler outer margins of the jet (2. in Fig. 2), where the volatile supersaturation and decompression rate of the magma.
rapid cooling of the pyroclast inhibited post-fragmentation vesicula- In the case of Hawaiian fountains, the maximum fountain height
tion processes. should theoretically be directly proportional to the degree of super-
saturation and nucleation rate in the shallow ascending magma and
5.2. Relationship between vesicle number density (Nmv) and fountain the N m v of the erupted tephra (Polacci et al., 2009; Stovall et al.,
height 2012).
In this study average Nmv for pyroclasts from each episode does not
High volatile supersaturations associated with disequilibrium correlate with the observed maximum fountain height (Fig. 10). In fact,
(rapid) magma ascent drives the rate of bubble nucleation, as dem- pyroclasts from the highest fountain (episode 32, 391 m) have the low-
onstrated by experiments (Gardner, 2007; Le Gall and Pichavant, est average Nmv (6.13 × 105 cm−3). Fig. 11 shows that the Nmv of the pri-
2016a, 2016b) and numerical models (Burgisser and Degruyter, mary pyroclasts from Episodes 37, 40, 44 and 45 may be more closely
2015; Gonnermann and Manga, 2013; Head and Wilson, 1989; related to fountain height than the average Nmv for each episode, and
Parfitt, 2004; Toramaru, 1995; Vergniolle and Gaudemer, 2015; implies that the microtextural data from primary pyroclasts better rep-
Vergniolle and Jaupart, 1990). Thus, bubble nucleation rates, and resents the vesiculation processes that were occurring within the shal-
therefore the Nmv of the erupted tephra, can be used as a proxy for low conduit prior to fragmentation.
30 S.J. Holt et al. / Journal of Volcanology and Geothermal Research 369 (2019) 21–34

episodes demonstrates that the magnitude of the change in VG/VL is


proportional to fountain height and thus residence time of the pyroclast
within the jet. Episodes 37 and 44 have maximum fountain heights of
352 and 308 m respectively, compared to 257 m for episode 45. We sug-
gest that lapilli remained above the glass transition temperature within
the relatively low jet of episode 45 for shorter times than during epi-
sodes 37 and 44. There was thus less time for post-fragmentation pro-
cesses to modify the vesicle population in episode 45 before
quenching, resulting in a large proportion of smaller vesicles, bubbly
textures and a negatively skewed polymodal density distribution
(Fig. 4).
Growth and coalescence of vesicles within the jet after
fragmentation results in an increase in average vesicle size, de-
crease in number density, and a thinning of vesicle walls, ulti-
mately forming closely packed polyhedral vesicles (Mangan and
Cashman, 1996). This vesicle evolution is reflected by the VSDs
for pyroclasts from each episode, most notably when comparing
episodes 45 and 37, which show the larger median and modal
vesicle sizes, coupled with fewer vesicles ≤0.16 mm in diameter,
with fountain height.

5.3.2. Episode 32
The Episode 32 pyroclasts show a simple trend in Fig. 9 with a similar
Nmv, yet large change in VG/VL. The high VG/VL pyroclasts 04B-16 and
04B-06 have foamy textures that are consistent with continued growth
Fig. 10. Plot of average Nmv versus fountain height for each eruptive episode. Black bars
of vesicles from bubbly textures (pyroclast 04B-20) in the jet after frag-
indicate the maximum and minimum Nmv. Numbers correspond to the episode.
mentation (Fig. 9C). The second lowest VG/VL pyroclast, 04B-38 has a
unique texture defined by a wide range in vesicle sizes but dominated
Of the primary pyroclasts found, pyroclast 02–04 from episode 44 by vesicles that are between 0.62 and 2.47 mm (Fig. 8). Large vesicles
and pyroclast 01–08 from episode 45 have quenched rinds and a vesicle are ragged or irregular in shape and separated by thick (100–400 μm)
size gradient from rind to interior (Fig. 11). These features suggest that, glass walls (Fig. 5D).
despite both pyroclasts having the lowest VG/VL values found for their Of the entire sample set described in this study, only pyroclast 04B-
respective episodes, even they underwent some post-fragmentation 38 from episode 32 has microlites; its glass, on average, contains 5–10%
vesicle modification. Therefore, episodes 32, 44 and 45 contain no microlites; some vesicle-poor regions contain up to 30% microlites
pyroclasts that truly exhibit pre-/syn- fragmentation microtextures, al- (Fig. 6). The qualitative textures of pyroclast 04B-38 suggest that it is,
though the texture of the quenched rims preserves some indications in fact, more evolved than both the foamy textures in the high VG/VL
of pre-/syn- eruptive conditions. Only pyroclasts 03–05 and 04A-05, pyroclasts (pyroclasts 04B–16 and 04–06) and the bubbly texture of
from the episodes 40 and 37 samples respectively, appear to be unmod- the other low VG/VL pyroclast 04B-20. Large, irregular vesicles separated
ified by post-fragmentation vesiculation processes and preserve the by thick glassy regions rich in microlites are indicative of magmas that
syn-fragmentation vesicle population. Both pyroclasts have bubbly tex- have undergone bubble coalescence, outgassing and microlite
tures in which larger and smaller vesicles define bands a few millime- crystallisation prior to quenching (e.g., Polacci et al., 2006; Sable et al.,
ters across (Fig. 11). While it is not possible to quantitatively describe 2009; Costantini et al., 2010). The source of this sub-population of
and compare the vesiculation histories and shallow conduit processes dense, texturally evolved pyroclasts is unknown (Figs. 4, 5d). Their tex-
of these high fountaining episodes based solely on these two pyroclasts, tures are commonly taken to reflect relatively long residence time be-
each of their qualitative textures suggest that the ascending melt is ther- fore eruption (e.g., Cimarelli et al., 2010; Sable et al., 2006). The repose
mally and (or) mechanically heterogeneous on a small scale during period before episode 32 was 38.4 days and shorter than the repose pe-
Hawaiian-style fountaining. riod before episode 40 (48.5 days), and yet there is no current textural
Due to the sparsity of primary textured pyroclasts, it is not possible, evidence that episode 40 produced observed pyroclasts with this tex-
based on Nmv, to quantitatively describe the fluid dynamics of the ture. There are at least two possible sources for these texturally evolved
magma in the shallow conduit in such a way that allows us to the pyroclasts: 1) magma that stalled in one of the subterranean reservoirs
make detailed comparisons between distinct fountaining episodes at a proposed by Garcia et al. (1992), or 2) lava that had previously ponded
single vent, or more importantly, between fountaining episodes at dif- and drained back in the conduit (Greenland et al., 1988). Whatever their
ferent vents or different volcanoes. origin, it is unclear why these ragged-textured pyroclasts are observed
only in the deposits of episode 32 and not in any of the other four
5.3. Relationship between vesicle-to-melt ratio (VG/VL) and fountain height episodes.

5.3.1. Episodes 37, 44 and 45 5.3.3. Episode 40


Episodes 37, 44 and 45 (Fig. 9A) pyroclasts show trends associated Episode 40 has the most complex vesiculation history of all five ep-
with continued growth and coalescence of vesicles after fragmentation, isodes. The VG/VL vs. Nmv plot suggests that, after fragmentation, some
most likely within the incandescent jet. For these episodes, continued pyroclasts underwent, bubble nucleation, growth, and coalescence
bubble growth and coalescence transforms primary bubbly textured (Fig. 9B). The growth plus coalescence trend is defined by pyroclasts
pyroclasts (low VG/VL, high Nmv) into mature foamy textured pyroclasts 03–05, 03–07 and 03–08. Pyroclasts 03–05 and 03–07 are primary
(high VG/VL, low Nmv). Episodes 37 and 44 pyroclasts both have large and have bubbly texture, while pyroclast 03–08 has foamy texture.
ranges in VG/VL (ranges of approximately 9 and 5, respectively), The trend defined by these two pyroclasts indicates that post-
whereas the range of VG/VL for episode 45 pyroclasts is relatively fragmentation vesicle growth and coalescence occurred within the in-
small (a change of approximately 3.5). The comparison of these candescent jet (similar to the processes described in Section 5.3.1).
S.J. Holt et al. / Journal of Volcanology and Geothermal Research 369 (2019) 21–34 31

Fig. 11. A) Plot of primary pyroclast Nmv versus fountain height for each eruptive episode for which primary pyroclasts were identified. B) Qualitative microtexture of primary pyroclasts
depicted in A).
32 S.J. Holt et al. / Journal of Volcanology and Geothermal Research 369 (2019) 21–34

Pyroclast 03–17 has a high Nmv value and moderate to high VG/VL Table 3
value and appears to define a trend of growth and nucleation. However, Calculated Δ(VG/VL) values for each eruption episode.

we believe this trend to be misleading as pyroclast 03–17 has separate Pyroclast Episode VG/VL Δ(VG/VL)a
zones of foamy texture and bubbly texture, and therefore two separate 01-33 Ep. 45 2.06 3.44
vesicle size populations. 01-08 2.50
01-06 3.76
01-01 5.49
02-04 Ep. 44 3.20 4.73
5.4. Quantitative microtextural signatures as a proxy for fountain height
02-08 5.54
02-05 7.93
We define the difference between the maximum and minimum 03-05 Ep. 40 2.86 7.50
VG/VL values obtained for pyroclasts from each eruption episode as 03-07 4.65
Δ(VG/VL). Calculated Δ(VG/VL) values are used to quantify the degree 03-17 7.20
03-08 10.36
of post-fragmentation vesicle growth that took place in order to 04A-05 Ep. 37 2.15 8.89
modify pyroclasts from bubbly texture to more mature foamy tex- 04A-09 2.94
ture. This can be used to infer the residence time of a pyroclast within 04A-01 7.85
the hot plume of the fountain above the glass transition temperature, 04A-04 11.05
04B-20 Ep. 32 1.64 7.57
which in turn, is dependent on the peak fountain height (Stovall
04B-38 2.41
et al., 2011; Stovall et al., 2012). 04B-16 7.55
Table 3 and Fig. 12 show Δ(VG/VL) values for Episodes 32, 37, 40, 04B-06 9.20
44 and 45 of the 1983–1986 Pu'u 'Ō'ō eruption as well as Δ(VG/VL) a
Change in vesicle to melt ratio.
values for Episodes 1, 15 and 16 of the 1959 Kīlauea Iki eruption
from (Stovall et al., 2011; Stovall et al., 2012), plotted against the
peak fountain height for each respective episode. This plot shows Acknowledgements
that, for all eight of these selected Hawaiian fountains there is a
good linear correlation between Δ(VG/VL) and peak fountain height. We sincerely thank scientists and staff at the Hawaiian Volcano Ob-
This correlation holds true whether for discrete fountaining episodes servatory for their assistance with this research. We thank Dr. Carolyn
from a single vent or two separate eruptions from different Parcheta for her detailed review of this manuscript. We thank the anon-
vents. This correlation suggests that Δ(V G/VL ) derived from the ymous reviewers for their constructive comments and suggestions in
microtextural analysis of multiple clasts from the same narrow earlier drafts of this manuscript. This research was supported by
stratigraphic interval could be used to ascertain the approximate
peak fountain height of unobserved Hawaiian-style fountaining
eruptions.
This methodology could potentially be applied to unobserved
fountaining eruptions at Kīlauea e.g. historical caldera eruptions
(Ferguson et al., 2016; May et al., 2015) to better understand its eruptive
history and therefore its associated hazards. Furthermore, by applying
this methodology to the deposits of Hawaiian fountaining at other ba-
saltic volcanoes around the world i.e. short duration, high-intensity
lava fountaining eruptions at Mount Etna Volcano, Italy (Ganci et al.,
2012; Polacci et al., 2009) it is possible to investigate the effects of com-
plex magma properties and conduit flow regimes on the textural rela-
tionships proposed in this study.

6. Conclusions

The five high Hawaiian fountaining episodes from the Pu'u 'Ō'ō erup-
tion presented in this study had similar vesicle populations before frag-
mentation, with bubbly textures and melt- corrected vesicle volume
density (Nmv) values between 2.2 × 106 cm−3 and 6.5 × 106 cm−3.
These Nmv values fill a gap between values previously recorded for
high Hawaiian fountains at Mauna Ulu and Kīlauea Iki (Parcheta et al.,
2013; Stovall et al., 2011; Stovall et al., 2012). However, the signature
of decompression rate and magma ascent is heavily overprinted by
post-fragmentation vesiculation processes. These processes result in a
reduction in the Nmv of tephra from these episodes, except in the most
primary pyroclasts defined by low VG/VL ratio and high Nmv. Growth
plus coalescence is the dominant post-fragmentation vesiculation
process that occurs within the jet of Hawaiian fountains. Post-
fragmentation vesiculation processes have a profound effect on the
vesicle microtextures to the extent that the theoretical relationships be-
tween Nmv, explosivity and fountain height in Hawaiian-style eruptions
does not hold true. We propose that the value range between the max-
Fig. 12. Plot showing Δ(VG/VL) vs peak fountain height for all fountaining episodes
imum and minimum vesicle-to-melt ratio (Δ(VG/VL)) for the tephra of sampled in this study, as well as three episodes (1, 15 and 16) from the 1969 Kīlauea Iki
Hawaiian fountaining eruptions can be used as a proxy for peak fountain eruption from Stovall et al. (2011) and Stovall et al. (2012). Linear regression is
height. calculated on all data points.
S.J. Holt et al. / Journal of Volcanology and Geothermal Research 369 (2019) 21–34 33

United States Geological Survey grant SV-ARRA-0004, Australian Houghton, B., Wilson, C., Del Carlo, P., Coltelli, M., Sable, J., Carey, R., 2004. The influence of
conduit processes on changes in style of basaltic Plinian eruptions: Tarawera 1886
Research Council grants DP110102196 and DE150101190 and National and Etna 122 BC. J. Volcanol. Geotherm. Res. 137 (1), 1–14.
Science Foundation, United States grants EAR-1131353 and EAR- Houghton, B., Taddeucci, J., Andronico, D., Gonnermann, H., Pistolesi, M., Patrick, M.R., Orr,
1521855. All SEM imaging was carried out at the Central Science T., Swanson, D., Edmonds, M., Gaudin, D., 2016. Stronger or longer: discriminating be-
tween Hawaiian and Strombolian eruption styles. Geology 44 (2), 163–166.
Laboratory at the University of Tasmania. Johnson, M.C., Anderson, A.T., Rutherford, M.J., 1994. Pre-eruptive volatile contents of
magmas. Rev. Mineral. Geochem. 30 (1), 281–330.
Appendix A. Sample sets were attributed to corresponding eruption Klug, C., Cashman, K., Bacon, C., 2002. Structure and physical characteristics of pumice
from the climactic eruption of Mount Mazama (Crater Lake), Oregon. Bull. Volcanol.
episodes using the following reference datasets 64 (7), 486–501.
Lautze, N.C., Houghton, B.F., 2005. Physical mingling of magma and complex eruption dy-
– Syn-eruption visual observations made by HVO staff members. namics in the shallow conduit at Stromboli volcano, Italy. Geology 33 (5), 425–428.
Lautze, N.C., Houghton, B.F., 2007. Linking variable explosion style and magma textures
– HVO eruption photograph archives.
during 2002 at Stromboli volcano, Italy. Bull. Volcanol. 69 (4), 445–460.
– National Oceanic and Atmospheric Administration (NOAA) Le Gall, N., Pichavant, M., 2016a. Experimental simulation of bubble nucleation and
historical windspeed and direction archives. magma ascent in basaltic systems: implications for Stromboli volcano. Am. Mineral.
101 (9), 1967–1985.
Le Gall, N., Pichavant, M., 2016b. Homogeneous bubble nucleation in H2O-and H2O-CO2-
bearing basaltic melts: results of high temperature decompression experiments.
J. Volcanol. Geotherm. Res. 327, 604–621.
Aerial photographic interpretation of tephra fall out by HVO staff Llewellin, E., Manga, M., 2005. Bubble suspension rheology and implications for conduit
members preceding each eruption episode. flow. J. Volcanol. Geotherm. Res. 143 (1), 205–217.
Manga, M., Castro, J., Cashman, K.V., Loewenberg, M., 1998. Rheology of bubble-bearing
magmas. J. Volcanol. Geotherm. Res. 87 (1), 15–28.
References Mangan, M.T., Cashman, K.V., 1996. The structure of basaltic scoria and reticulite and in-
ferences for vesiculation, foam formation, and fragmentation in lava fountains.
Abràmoff, M.D., Magalhães, P.J., Ram, S.J., 2004. Image processing with ImageJ. Biophoton. J. Volcanol. Geotherm. Res. 73 (1), 1–18.
Int. 11 (7), 36–42. May, M., Carey, R.J., Swanson, D.A., Houghton, B.F., 2015. Reticulite-producing fountains
Adams, N.K., Houghton, B.F., Hildreth, W., 2006. Abrupt transitions during sustained ex- from ring fractures in Kīlauea Caldera ca. 1500 CE. In: Carey, R., Valérie, C., Poland,
plosive eruptions: examples from the 1912 eruption of Novarupta, Alaska. Bull. M., Weis, D. (Eds.), Hawaiian Volcanoes: From Source to Surface. John Wiley &
Volcanol. 69 (2), 189–206. Sons, pp. 351–367.
Burgisser, A., Degruyter, W., 2015. Magma ascent and degassing at shallow levels. The En- Moore, R.B., Helz, R.T., Dzurisin, D., Eaton, G.P., Koyanagi, R.Y., Lipman, P.W., Lockwood,
cyclopedia of Volcanoes, Second edition Elsevier, pp. 225–236. J.P., Puniwai, G.S., 1980. The 1977 eruption of Kilauea volcano, Hawaii. J. Volcanol.
Carey, R.J., Manga, M., Degruyter, W., Swanson, D., Houghton, B., Orr, T., Patrick, M., 2012. Geotherm. Res. 7 (3), 189–210.
Externally triggered renewed bubble nucleation in basaltic magma. The October 12 Papale, P., 1998. Volcanic conduit dynamics. In: Freundt, A., Rosi, M. (Eds.), From Magma
2008 eruption at Halema 'uma 'u Overlook vent, Kīlauea, Hawai'i, USA. to Tephra: Modelling Physical Processes of Explosive Volcanic Eruptions. Elsevier Sci-
Cashman, K.V., Mangan, M.T., 1994. Physical aspects of magmatic degassing; II, Con- ence, pp. 55–91.
straints on vesiculation processes from textural studies of eruptive products. Rev. Papale, P., 2005. Determination of total H2O and CO2 budgets in evolving magmas from
Mineral. Geochem. 30 (1), 447–478. melt inclusion data. J. Geophys. Res. Solid Earth 110 (B3) (1978–2012).
Cimarelli, C., Di Traglia, F., Taddeucci, J., 2010. Basaltic scoria textures from a zoned con- Parcheta, C., Houghton, B., Swanson, D., 2013. Contrasting patterns of vesiculation in low,
duit as precursors to violent Strombolian activity. Geology 38 (5), 439–442. intermediate, and high Hawaiian fountains: a case study of the 1969 Mauna Ulu
Costantini, L., Houghton, B., Bonadonna, C., 2010. Constraints on eruption dynamics of ba- eruption. J. Volcanol. Geotherm. Res. 255, 79–89.
saltic explosive activity derived from chemical and microtextural study: the example Parfitt, E.A., 2004. A discussion of the mechanisms of explosive basaltic eruptions.
of the Fontana Lapilli Plinian eruption, Nicaragua. J. Volcanol. Geotherm. Res. 189 (3), J. Volcanol. Geotherm. Res. 134 (1), 77–107.
207–224. Parfitt, E., Wilson, L., 1995. Explosive volcanic eruptions—IX. The transition between
Edmonds, M., Gerlach, T.M., 2007. Vapor segregation and loss in basaltic melts. Geology Hawaiian-style lava fountaining and Strombolian explosive activity. Geophys. J. Int.
35 (8), 751–754. 121 (1), 226–232.
Fedotov, S., Chirkov, A., Gusev, N., Kovalev, G., Slezin, Y.B., 1980. The large fissure eruption Polacci, M., 2005. Constraining the dynamics of volcanic eruptions by characterization of
in the region of Plosky Tolbachik volcano in Kamchatka, 1975–1976. Bull. Volcanol. pumice textures. Annals of Geophysics.
43 (1), 47–60. Polacci, M., Corsaro, R.A., Andronico, D., 2006. Coupled textural and compositional charac-
Ferguson, D.J., Gonnermann, H.M., Ruprecht, P., Plank, T., Hauri, E.H., Houghton, B.F., terization of basaltic scoria: insights into the transition from Strombolian to fire foun-
Swanson, D.A., 2016. Magma decompression rates during explosive eruptions of Kī- tain activity at Mount Etna, Italy. Geology 34 (3), 201–204.
lauea volcano, Hawaii, recorded by melt embayments. Bull. Volcanol. 78 (10), 71. Polacci, M., Burton, M.R., La Spina, A., Murè, F., Favretto, S., Zanini, F., 2009. The role of syn-
Ganci, G., Harris, A.J., Del Negro, C., Guéhenneux, Y., Cappello, A., Labazuy, P., Calvari, S., eruptive vesiculation on explosive basaltic activity at Mt. Etna, Italy. J. Volcanol.
Gouhier, M., 2012. A year of lava fountaining at Etna: volumes from SEVIRI. Geophys. Geotherm. Res. 179 (3), 265–269.
Res. Lett. 39 (6). Porritt, L.A., Russell, J.K., Quane, S.L., 2012. Pele's tears and spheres: examples from Kilauea
Garcia, M.O., Rhodes, J.M., Wolfe, E., Ulrich, G.E., Ho, R., 1992. Petrology of lavas from ep- Iki. Earth Planet. Sci. Lett. 333–334, 171–180.
isodes 2–47 of the Puu Oo eruption of Kilauea Volcano, Hawaii: evaluation of mag- Richter, D.H., Eaton, J., Murata, K., Ault, W., Krivoy, H., 1970. Chronological Narrative of the
matic processes. Bull. Volcanol. 55 (1–2), 1–16. 1959–60 Eruption of Kilauea Volcano, Hawaii. pp. 2330–7102.
Gardner, J.E., 2007. Heterogeneous bubble nucleation in highly viscous silicate melts dur- Rust, A., Cashman, K., 2011. Permeability controls on expansion and size distributions of
ing instantaneous decompression from high pressure. Chem. Geol. 236 (1), 1–12. pyroclasts. J. Geophys. Res. Solid Earth 116 (B11).
Gardner, J.E., Thomas, R.M., Jaupart, C., Tait, S., 1996. Fragmentation of magma during Ryan, M.P., 1988. The mechanics and three-dimensional internal structure of active mag-
Plinian volcanic eruptions. Bull. Volcanol. 58 (2–3), 144–162. matic systems: Kilauea Volcano, Hawaii. J. Geophys. Res. Solid Earth 93 (B5),
Gonnermann, H.M., Manga, M., 2007. The fluid mechanics inside a volcano. Annu. Rev. 4213–4248 (1978–2012).
Fluid Mech. 39, 321–356. Sable, J.E., Houghton, B.F., Del Carlo, P., Coltelli, M., 2006. Changing conditions of magma
Gonnermann, H.M., Manga, M., 2013. Dynamics of magma ascent in the volcanic conduit. ascent and fragmentation during the Etna 122 BC basaltic Plinian eruption: Evidence
In: Fagents, S., Gregg, T.K., Lopes, R.M. (Eds.), Modeling Volcanic Processes: the Phys- from clast microtextures. J. Volcanol. Geotherm. Res. 158 (3), 333–354.
ics and Mathematics of Volcanism. Cambridge University Press. Sable, J., Houghton, B., Wilson, C., Carey, R., 2009. Eruption mechanisms during the climax
Greenland, L., Okamura, A., Stokes, J., 1988. Constraints on the mechanics of eruption. In: of the Tarawera 1886 basaltic Plinian eruption inferred from microtextural character-
Wolfe, E. (Ed.), The Puu O eruption of Kilauea Volcano, Episodes 1–20, January 3, istics of the deposits. Studies in Volcanology: The Legacy of George Walker. 2. Spec.
1983, to June 8, 1984. U.S. Government Printing Office. Publ. IAVCEI, pp. 129–154.
Head, J.W., Wilson, L., 1989. Basaltic pyroclastic eruptions: influence of gas-release pat- Sahagian, D.L., Proussevitch, A.A., 1998. 3D particle size distributions from 2D observa-
terns and volume fluxes on fountain structure, and the formation of cinder cones, tions: stereology for natural applications. J. Volcanol. Geotherm. Res. 84 (3), 173–196.
spatter cones, rootless flows, lava ponds and lava flows. J. Volcanol. Geotherm. Res. Shea, T., Houghton, B.F., Gurioli, L., Cashman, K.V., Hammer, J.E., Hobden, B.J., 2010. Tex-
37 (3), 261–271. tural studies of vesicles in volcanic rocks: an integrated methodology. J. Volcanol.
Heliker, C., Swanson, D.A., Takahashi, T.J., 2003. The Pu'u 'Ō'ō-Kūpaianaha Eruption of Kī- Geotherm. Res. 190 (3), 271–289.
lauea Volcano, Hawai'i: The First 20 Years. 1676. US Dept. of the Interior, US Geolog- Spilliaert, N., Allard, P., Métrich, N., Sobolev, A., 2006. Melt inclusion record of the condi-
ical Survey. tions of ascent, degassing, and extrusion of volatile-rich alkali basalt during the pow-
Houghton, B., Gonnermann, H., 2008. Basaltic explosive volcanism: Constraints from de- erful 2002 flank eruption of Mount Etna (Italy). J. Geophys. Res. 111, B04203.
posits and models. Chem. Erde-Geochem. 68 (2), 117–140. Stovall, W.K., Houghton, B., Gonnermann, H., Fagents, S., Swanson, D., 2011. Eruption dy-
Houghton, B., Wilson, C., 1989. A vesicularity index for pyroclastic deposits. Bull. Volcanol. namics of Hawaiian-style fountains: the case study of episode 1 of the Kīlauea Iki
51 (6), 451–462. 1959 eruption. Bull. Volcanol. 73 (5), 511–529.
Houghton, B., Wilson, C., Smith, I., 1999. Shallow-seated controls on styles of explosive ba- Stovall, W.K., Houghton, B.F., Hammer, J.E., Fagents, S.A., Swanson, D.A., 2012. Vesiculation
saltic volcanism: a case study from New Zealand. J. Volcanol. Geotherm. Res. 91 (1), of high fountaining Hawaiian eruptions: episodes 15 and 16 of 1959 Kīlauea Iki. Bull.
97–120. Volcanol. 74 (2), 441–455.
34 S.J. Holt et al. / Journal of Volcanology and Geothermal Research 369 (2019) 21–34

Swanson, D.A., Duffield, W.A., Jackson, D.B., Peterson, D.W., 1979. Chronological Narrative Wilson, L., 1980. Relationships between pressure, volatile content and ejecta velocity in
of the 1969–71 Mauna Ulu Eruption of Kilauea volcano, Hawaii. US Govt. Print. Off., three types of volcanic explosion. J. Volcanol. Geotherm. Res. 8 (2), 297–313.
pp. 2330–7102. Wilson, L., Head, J., 1981. Ascent and eruption of basaltic magma on the Earth and Moon.
Taddeucci, J., Edmonds, M., Houghton, B.F., James, M., Vergniolle, S., 2015. Hawaiian and J. Geophys. Res. 86 (B4), 2971–3001.
Strombolian eruptions. In: Sigurdsson, H. (Ed.), Encyclopedia of Volcanoes. Elsevier. Wolfe, E., Garcia, M., Jackson, D., Koyanagi, R., Neal, C., Okamura, A., 1987. The Puu O erup-
Toramaru, A., 1995. Numerical study of nucleation and growth of bubbles in viscous tion of Kilauea Volcano, Episodes 1–20, January 3, 1983, to June 8, 1984. In: Decker, R.,
magmas. J. Geophys. Res. Solid Earth 100 (B2), 1913–1931. Wright, T., Stauffer, P. (Eds.), Volcanism in Hawaii. 1. U.S. Government Printing
Vergniolle, S., Gaudemer, Y., 2015. From reservoirs and conduits to the surface: Review of Office.
role of bubbles in driving basaltic eruptions. Hawaiian Volcanism: From Source to Wolfe, E., Neal, C., Banks, N., Duggan, T., 1988. Geologic observations and chronology of
Surface. 208. Am. Geophysical Union Monograph, pp. 289–322. eruptive events. US Geol. Surv. Prof. Pap. 1463, 1–97.
Vergniolle, S., Jaupart, C., 1990. Dynamics of degassing at Kilauea volcano, Hawaii. Wolff, J., Sumner, J., 2000. Lava fountains and their products. In: Sigurdsson, H., Houghton,
J. Geophys. Res. 95 (B3), 2793–2809. B.F., SR, McNutt, Rymer, H., Stix, J. (Eds.), Encyclopedia of Volcanoes. Academic Press,
Walker, G.P., 1973. Explosive volcanic eruptions—a new classification scheme. Geol. San Diego. Academic Press, San Diego, pp. 321–329.
Rundsch. 62 (2), 431–446.

You might also like